06.05.2013 Views

Alkyl Ketene Dimer (AKD) sizing – a review

Alkyl Ketene Dimer (AKD) sizing – a review

Alkyl Ketene Dimer (AKD) sizing – a review

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Alkyl</strong> <strong>Ketene</strong> <strong>Dimer</strong> (<strong>AKD</strong>) <strong>sizing</strong> <strong>–</strong> a <strong>review</strong><br />

Tom Lindström and Per Tomas Larsson, STFI-Packforsk AB, Stockholm, Sweden<br />

KEYWORDS: <strong>Alkyl</strong> ketene dimer, Internal <strong>sizing</strong>,<br />

Hydrophobicity, Retention, Spreading Reaction, Mechanisms,<br />

Paper, Board, Manufacture, Review<br />

SUMMARY: Over the years, there have been great efforts to<br />

try to develop cellulose reactive <strong>sizing</strong> agents. The assumption<br />

in these developments have been that the covalent linkage<br />

allows permanent attachment of hydrophobic groups in a highly<br />

oriented state, which makes <strong>sizing</strong> possible at very low levels of<br />

added chemical. The main requirement of the molecule is that it<br />

should have a balance between the reactivity towards water,<br />

because of the necessity of making stable emulsions or dispersions,<br />

and its reactivity towards cellulose. These assumptions<br />

are to some extent mutually exclusive and a compromise must<br />

be sought. Although, many different types have been tried out<br />

over the years the most important sizes used are the <strong>Alkyl</strong><br />

<strong>Ketene</strong> <strong>Dimer</strong>s (<strong>AKD</strong>) and the Alkenyl Succinic Anhydrides<br />

(ASA). These <strong>sizing</strong> agents are at the opposite in terms of stability<br />

of hydrolysis and reactivity towards cellulose, where <strong>AKD</strong>s<br />

are the least reactive species and fairly stable towards hydrolysis,<br />

whereas ASAs are very reactive towards cellulose, but also<br />

sensitive to hydrolysis. The mechanism of action is fairly well<br />

known for <strong>AKD</strong>, but less known for ASA and <strong>AKD</strong>-<strong>sizing</strong> can<br />

be regarded as a pretty mature field from a scientific point of<br />

view. The aim of this contribution is to summarize the fundamental<br />

features of <strong>AKD</strong>-<strong>sizing</strong> and discuss and highlight the<br />

most important aspects for the practical papermaker.<br />

Over the years there have been many <strong>review</strong>s (e.g. (Dumas<br />

1975; Reynolds 1989; Eklund and Lindström 1991; Hodgson<br />

1994; Roberts 1997; Hubbe 2006)) in the field of <strong>AKD</strong>-<strong>sizing</strong>,<br />

but there have been extensive recent research activities over the<br />

past 10 years and there is a need for a comprehension of these<br />

research activities.<br />

ADDRESS OF THE AUTHORS: Tom Lindström<br />

(tom.lindstrom@stfi.se) and Per Tomas Larsson<br />

(tomas.larsson@stfi.se): STFI-Packforsk AB, Box 5604,<br />

SE-114 86 Stockholm, Sweden.<br />

Corresponding author: Tom Lindström<br />

Basic Chemical Features of <strong>AKD</strong><br />

<strong>Alkyl</strong> ketene dimers were the results of direct development<br />

efforts in the late 40´s (Downey 1949). These<br />

investigations demonstrated that the parent molecule, the<br />

diketene could derivatize hydroxyl groups and in particular<br />

those of cellulose. The strained lactone ring in<br />

ketene dimers can react both with cellulose and water<br />

forming either the β-keto ester or the β-keto acid, which<br />

spontaneously decarboxylates to the corresponding ketone<br />

as shown in Fig 1. The ketone is incapable of reacting<br />

with cellulose. The balance of reaction with cellulose and<br />

hydrolysis is subtle, as with all reactive sizes, but the<br />

reaction is favoured and, hence, <strong>AKD</strong> can be used as a<br />

<strong>sizing</strong> agent under commercial papermaking conditions.<br />

The nucleophilic reaction with cellulose can also be accelerated<br />

with various so-called promoters, which will be<br />

discussed below.<br />

202 Nordic Pulp and Paper Research Journal Vol 23 no. 2/2008<br />

Fig 1. <strong>Alkyl</strong> ketene dimers can react either with cellulose forming the β-keto ester<br />

or with water forming the β-keto acid, which spontaneously decarboxylates forming<br />

the corresponding ketone.<br />

Commercial <strong>AKD</strong>s are prepared from long fatty acids via<br />

their acid chlorides, which then dimerize to the corresponding<br />

alkyl ketene dimer (Fig 2).<br />

Fig 2. Formation of alkyl ketene dimers from the corresponding fatty acid chlorides.<br />

The linear saturated <strong>AKD</strong>s are waxy substances, water<br />

insoluble solids with melting points around 50°C, being<br />

mixtures of C-14 to C-18 fatty acids when manufactured<br />

from commercially available fatty acids. C-16 is the predominant<br />

fatty acid in the most used formulations. The<br />

efficiency increases with carbon chain length from C-8<br />

and levels off at C-20 (Brungardt and Varnell 2005).<br />

It is also known that small amounts (usually less than<br />

10% (Hardell and Woodbury 2002) of <strong>AKD</strong>-oligomers<br />

are present in <strong>AKD</strong> (Bottorff 1993; Bottorff 1994;<br />

Asakura et al. 2006a) and that these oligomers give<br />

relatively poor <strong>sizing</strong>.<br />

Alkenylsubstituted <strong>AKD</strong>s (emulsions = liquid in liquid<br />

phase) are also commercially available and they have<br />

been found to give better runnability on high speed<br />

handling equipment than paper sized with suspensions<br />

(solid phase in a liquid phase) (Brungardt and Gast<br />

1996), but they are slightly inferior to alkyl ketene dimers<br />

(Isogai and Asakura 2001a, 2001b).<br />

Emulsification<br />

<strong>AKD</strong>s are dispersed using high-pressure homogenizers at<br />

elevated temperatures. The most frequently used<br />

stabilizers are cationic starches in conjunction with<br />

lignosulfonates/naphthalene sulfonic acids. Waxy maize<br />

starches with no propensity to retrogradation are the<br />

preferred choice of starch. Most commercially used<br />

dispersions (note: both emulsions and suspensions are dispersions)<br />

are therefore amphoteric (Isogai 1997b), which<br />

is important from a retention point of view (see below).


It is important to avoid surface active substances in the<br />

dispersion formulation, because they may interfere with<br />

<strong>sizing</strong> (see below). The dispersions are usually made<br />

slightly cationic in order for them to have a natural<br />

substantivity to negatively charged fibres, but anionic<br />

dispersions are also used in commercial practice. In order<br />

to avoid hydrolysis of the <strong>AKD</strong>, the pH is kept around 3<br />

in the formulations.<br />

There have been some recent studies on the stability of<br />

<strong>AKD</strong>-dispersions. It has been found that the stability<br />

increases by the presence of <strong>AKD</strong>-oligomers (Asakura et<br />

al. 2006a) and that fatty acid anhydrides (byproduct in<br />

the <strong>AKD</strong>-dispersion) decrease the heat shock stability<br />

(Asakura et al. 2006b). The effect of various colloidal<br />

substances present in process waters have also been investigated<br />

(Mattsson et al. 2001). <strong>AKD</strong> dispersions basically<br />

behave as electrostatically stabilized colloids.<br />

Consecutive events of <strong>AKD</strong>-<strong>sizing</strong><br />

The consecutive events in <strong>AKD</strong>-<strong>sizing</strong> are (Lindström<br />

and Söderberg 1986a):<br />

· Retaining the <strong>AKD</strong>-size using appropriate retention<br />

strategies for the size<br />

· Spreading/size migration to a monolayer<br />

· Chemical reaction of the size with the cellulosic fibres<br />

The retention mechanism is, in theory, heterocoagulation,<br />

where cationic size particles are attached to the negatively<br />

charged fibres. This is expected to give a good<br />

distribution of the dispersion particles, but practice shows<br />

that size distribution is not critical, because of extensive<br />

spreading on the fibre surfaces. More important is that<br />

effective retention aids must be used for the purpose.<br />

Heterocoagulation, cannot be used to retain <strong>AKD</strong>-particles<br />

under high fibre-fibre shear conditions. A high<br />

single pass retention is important, because recirculated<br />

size is hydrolyzed in the white water of a paper machine.<br />

When the size particles have been deposited on the<br />

fibres, the reaction with cellulose is insignificant because<br />

very few molecules are in molecular contact with cellulose.<br />

Because of the high surface tension of water, no spreading<br />

can take place until the water has been removed<br />

and the size particles are in direct contact with air.<br />

Hence, an air-<strong>AKD</strong> surface must be formed before<br />

spreading can take place and this takes place during<br />

drying at a solids content exceeding 60%. The spreading<br />

continues until a monomolecular layer has been formed.<br />

This layer then reacts with the hydroxyl groups of<br />

cellulose. The reaction is slow at low pH-values and, in<br />

practice; <strong>AKD</strong> cannot be used except in the neutral or<br />

slightly alkaline pH-range. Moreover, <strong>sizing</strong> accelerators<br />

(most commonly HCO -<br />

3) are almost always required for<br />

effective development of <strong>sizing</strong>.<br />

Retention of <strong>AKD</strong><br />

Cationic <strong>AKD</strong>-particles are theoretically retained by a<br />

heterocoagulation deposition mechanism onto the negatively<br />

charged fibres, but studies (Johansson and<br />

Lindström 2004b) have shown that the electrostatic<br />

attraction is screened by electrolyte concentrations<br />

commonly present in process waters, so retention aids<br />

must always be used in commercial practice. Secondly,<br />

the deposition is a highly dynamic process, where <strong>AKD</strong>particles<br />

are rapidly deposited, after which they are<br />

sheared off the fibre surfaces (Lindström and Söderberg<br />

1986c; Champ and Ettl 2004).<br />

The surface charge of fibres is important for the selfretention<br />

(retention of <strong>AKD</strong> without retention aid<br />

addition) of <strong>AKD</strong>, as self-retention is a heterocoagulation<br />

mechanism (Isogai et al. 1997; Lindström and Glad-<br />

Nordmark 2007a). There is an optimum surface charge<br />

density of fibres and an optimum electrolyte concentration<br />

for maximum <strong>AKD</strong>-retention (Lindström and Glad-<br />

Nordmark 2007a). These studies are, however, laboratory<br />

studies, and in commercial practice self-retention is most<br />

likely negligible due to high electrolyte concentrations<br />

and the high shear to which the fibre suspension is<br />

subjected.<br />

The retention of <strong>AKD</strong> by cationic polyelectrolytes has<br />

been studied by several other groups (Esser and Ettl<br />

1997; Hasegawa et al. 1997; Isogai 1997a, 1997b). It is<br />

not self-evident that it should be possible to retain<br />

cationic <strong>AKD</strong>-dispersions using cationic polyelectrolytes.<br />

Due to the fact that <strong>AKD</strong>-dispersions usually are<br />

dispersed using a combination of lignosulfonates and<br />

cationic starch, the net cationically charged dispersion<br />

particles have in fact an amphoteric nature. Hence<br />

cationic polyelectrolytes can interact with the negative<br />

sites on the dispersion particles and retain the <strong>AKD</strong>-particles.<br />

As shown in Fig 3, anionic <strong>AKD</strong>-particles are easier<br />

to retain by cationic polyelectrolytes than cationic <strong>AKD</strong>particles<br />

(Johansson and Lindström 2004b). The<br />

appropriate choice of an efficient retention aid system is<br />

crucial, because <strong>AKD</strong> will be subject to hydrolysis when<br />

circulated in the short circulation of a paper mill, as will<br />

be discussed below.<br />

The use of hydrophobic retention aids has also been<br />

practised (Riebeling et al. 1996, 1999) to treat filler so<br />

that the interaction between filler particles and <strong>AKD</strong><br />

could be minimized. Obviously <strong>AKD</strong> cannot react with<br />

fillers and CaCO 3-based fillers promote hydrolysis (see<br />

below).<br />

Pre-flocculation of <strong>AKD</strong> leads to higher <strong>AKD</strong>-retention<br />

(Mattsson et al. 2002), which in itself is beneficial to<br />

<strong>sizing</strong>. Agglomeration is not critical to <strong>sizing</strong> (Johansson<br />

and Lindström 2004b); neither is the particle size<br />

(Petander et al. 1998) because <strong>AKD</strong> spreads over the<br />

fibre surfaces anyhow.<br />

In the practical use of <strong>AKD</strong>, the point of addition is of<br />

course dependent on the wet end system and the addition<br />

order of other chemical adjuvants. As a rule of thumb it is<br />

advantageous to add <strong>AKD</strong> to the high consistency stock<br />

only a short time before dilution takes place in the short<br />

circulation. The <strong>AKD</strong>-particles are deposited at a rapid<br />

rate at high consistency but are subsequently sheared off<br />

the fibre surfaces by the other fibres in a stirred pulp<br />

suspension (Lindström and Söderberg 1986a; Champ and<br />

Ettl 2004). Both high shear and long contact times are<br />

known to reduce <strong>AKD</strong> retention. As a rule, both <strong>AKD</strong>,<br />

Nordic Pulp and Paper Research Journal Vol 23 no. 2/2008 203


Fig 3. a) Retention of cationic <strong>AKD</strong>-particles onto bleached kraft pulps using various cationic polyelectrolytes. b) Retention of anionic <strong>AKD</strong>-particles onto bleached kraft pulp<br />

using various cationic polyelectrolytes (Johansson and Lindström 2003b).<br />

fines and fillers follow the general retention level and<br />

there is little selective retention of certain types of dispersed<br />

material. Anionic dissolved substances in the stock,<br />

such as hemicelluloses and lignin residues, are generally<br />

detrimental to size retention (Lindström and Söderberg<br />

1986c).<br />

Spreading/size migration<br />

The distribution of <strong>AKD</strong>-size on the fibres occurs in the<br />

drying section as discussed above. <strong>AKD</strong> readily spreads<br />

on cellulose, because the cellulose surface is a high-energy<br />

surface. The free energy of spreading, ∆G s of <strong>AKD</strong> on<br />

a cellulose surface can be written:<br />

∆G s = γ(cellulose/<strong>AKD</strong>) + γ(<strong>AKD</strong>) - γ (cellulose) [1]<br />

The surface free energy of <strong>AKD</strong> is 27 mJ/m 2 (Garnier<br />

and Godbout 2000) and the surface free energy of dry<br />

cellulose has been determined to about 57 mJ/m 2 .<br />

(Lundqvist and Ödberg 1997; Luner and Oh 2001). If<br />

γ(cellulose/<strong>AKD</strong>) is small, the conditions for spreading,<br />

∆G s < 0, are fulfilled. For an <strong>AKD</strong> particle trapped in<br />

between a fibre-fibre bond the free energy of spreading,<br />

∆G s = 2γ(cellulose/<strong>AKD</strong>), which is a positive quantity,<br />

because it is associated with the cleavage of a high<br />

energy surface. Hence, <strong>AKD</strong>-particles trapped in between<br />

fibre-fibre bonds cannot spread and cannot react with<br />

cellulose. This phenomenon is the hypothesis for the<br />

limited reaction with cellulose (Lindström and O´Brian<br />

1986b). Spreading has been manifested by several investigators<br />

(Roberts and Garner 1985; Roberts et al. 1985;<br />

Ödberg et al. 1987; Seppänen et al. 2000; Horn 2001;<br />

Shchukarev et al. 2003). The spreading of <strong>AKD</strong> on<br />

cellulose should, however, not be associated with the<br />

common hydrodynamic phenomena of spreading<br />

(Cazabat 1989), which is a very rapid process. Instead, it<br />

has been suggested that spreading takes place by the<br />

surface diffusion of an autophobic monolayer of <strong>AKD</strong> on<br />

cellulose (Seppänen et al. 2000), which is a slower<br />

phenomena than hydrodynamic spreading. The apparent<br />

surface diffusion coefficient of <strong>AKD</strong> on cellulose have<br />

been calculated to around 10 -11 m 2 /s at 50-80°C<br />

204 Nordic Pulp and Paper Research Journal Vol 23 no. 2/2008<br />

(Seppänen et al. 2000; Shchukarev et al. 2003). As <strong>AKD</strong><strong>sizing</strong><br />

particles typically have the dimension of the order<br />

of a micrometer, such a droplet would on a cellulose<br />

surface spread within 10 sec, using this diffusion<br />

coefficient. The time of reaction is typically of the order<br />

of at least 5 minutes, hence spreading/surface diffusion is<br />

not the rate-determining step in <strong>sizing</strong>.<br />

More recent investigations have, however, challenged<br />

the traditional spreading view. Thus, Garnier et al<br />

(Garnier et al. 1998, 1999; Garnier and Godbout 2000)<br />

find that <strong>AKD</strong> only wets, but not spreads, on cellulose<br />

and claim vapour phase <strong>sizing</strong> as a <strong>sizing</strong> mechanism for<br />

<strong>AKD</strong>. Later investigations show vapour phase type of<br />

<strong>sizing</strong> with ASA but not with <strong>AKD</strong> at drying temperatures<br />

below 100°C. (Yu and Garnier 2002). A capillary<br />

wicking mechanism is also suggested by Garnier and<br />

Godbout (2000). One possibility is that the reservoir in<br />

the de Gennes type of experiments conducted by Garnier<br />

was simply too small. In this type of experiment a drop of<br />

the <strong>sizing</strong> agent is applied to a thread (cellulosic) and the<br />

contact angle is observed. A pre-requisite is that there is a<br />

sufficient surface area available for spreading to occur,<br />

otherwise spreading will stop when the available surface<br />

area is saturated with the monomolecular layer of the<br />

<strong>sizing</strong> agent, spreading stops and a finite contact angle is<br />

observed.<br />

Shen et al. also in a number of papers (Shen et al.<br />

2001a, 2001b; Shen and Parker 2003) claim that the<br />

classical view has been proven wrong and advocate<br />

mechanisms along the lines of Garnier and co-workers.<br />

These authors also claim there is no reaction between<br />

cellulose and <strong>AKD</strong>. Later investigations from this group<br />

(Hutton and Chen 2004), however, also show vapour<br />

phase <strong>sizing</strong> to be an insignificant phenomenon in<br />

practical papermaking <strong>sizing</strong>. The group has now (Shen<br />

and Parker 2003; Shen et al. 2005) adopted the autophobic<br />

precursor mechanism suggested by Seppänen et<br />

al. (2000).<br />

In our laboratory, a simple experiment was conducted<br />

to show that spreading does not take place in the gasphase<br />

and that spreading easily takes place and over<br />

macroscopic dimensions. Basically, two sets of experi-


ments were conducted in additon to a reference sheet,<br />

# 1 (Table 1). In the first experiment (# 2, Table 1) a sized<br />

sheet was couched (after wet-pressing) between two<br />

unsized sheets. This pack of sheets was then dried at<br />

90ºC and the size retention and extent of reaction was<br />

determined using radioactive labelled <strong>AKD</strong> as described<br />

in our previous publications on <strong>AKD</strong>, e.g., Lindström and<br />

Söderberg (1986a). After drying the stack of sheet was<br />

simply delaminated and the total and reacted amount of<br />

<strong>AKD</strong> could be determined. In the second experiment<br />

(# 3, Table 1) a thin polytetrafluoroethylene (PTFE) wire<br />

(allowing gas-phase transfer through the wire) was<br />

placed between the sheets before drying.<br />

These experiments showed two things: <strong>AKD</strong> could<br />

spread across the whole stack of sheets, although the<br />

mid-sheet still had the highest amount of <strong>AKD</strong>. In the<br />

experiment using the PTFE wire, no transfer between the<br />

sheets took place and consequently there was no gasphase<br />

spreading across the pile of sheets.<br />

Table 1.Results from <strong>AKD</strong>-transfer experiments. <strong>AKD</strong>-sized sheets (2.2 and 3.2)<br />

were stacked between unsized sheets (sheets: 2.1, 2.3 and 3.1, 3.3) after wet<br />

pressing and subsequently dried at 90ºC (series # 2) for 10 min. After drying, the<br />

sheets were post cured at 110ºC for 10 min. The reference sheet was dried at<br />

90ºC for 10 min and then at 110ºC for 10 min. Experiment # 3 was conducted as<br />

2.1-2.3 but a thin polytetrafluoroethylene wire was inserted between the sheets<br />

before couching to prevent size migration but allow for possible gas phase transfer<br />

of <strong>AKD</strong>. 0.1% C-PAM (D.S. = 20 mole % cationic groups) was used as a<br />

retention aid in the experiments. n.d. = not determined. Data not published previously.<br />

Sheet Content <strong>AKD</strong> <strong>AKD</strong> <strong>AKD</strong><br />

in sample retention reacted amount<br />

Exp. #/Sheet type mg/g % mg/g %<br />

1.1 <strong>AKD</strong> (1.5 mg/g) 1.15 76.7 0.46 40.0<br />

(Reference) +C-PAM<br />

2.1 no <strong>AKD</strong> 0.16 0.07 43.8<br />

2.2 <strong>AKD</strong> (1.5 mg/g) 0.88 80.0 0.38 43.2<br />

+C-PAM<br />

2.3 no <strong>AKD</strong> 0.16 0.06 37.5<br />

3.1 no <strong>AKD</strong> [<strong>AKD</strong>] and the equation reads:<br />

dx<br />

dt<br />

( )<br />

= K [<strong>AKD</strong>] = K [<strong>AKD</strong>] = K a− x<br />

[3]<br />

av<br />

1 2<br />

where a is the maximum amount of reacted <strong>AKD</strong> and<br />

consequently [<strong>AKD</strong>] av, the available amount for reaction<br />

should be = a <strong>–</strong> x. Integration from t = 0 to t leads to:<br />

ln a<br />

a x Kt = [4]<br />

−<br />

Thus, if the quantity ln(a/(a <strong>–</strong> x)) is plotted versus time,<br />

straight lines should be obtained and the reaction follows<br />

what could be called a pseudo first order reaction. This is<br />

illustrated in Fig 4a, which shows such graphs for the<br />

reaction of <strong>AKD</strong> onto a bleached kraft pulp at different<br />

pH-values. It is also shown that the straight lines intersect<br />

the x-axis at a common point in time, independent of pH.<br />

This is the time interval required for drying of the sheet<br />

to a sufficiently high solids content for the <strong>AKD</strong> to begin<br />

spreading. The spreading reaction as such is much faster<br />

than the drying stage, which is independent of the pHvalue<br />

as the rate of drying is independent on pH for this<br />

type of pulp. If the same experiment then is performed at<br />

different pH-values, the corresponding values between<br />

the reaction rate constant K at different pH-values and<br />

temperatures can be obtained and the Arrhenius activation<br />

energies for the <strong>AKD</strong>/cellulose reaction can be<br />

calculated from the graphs in Fig 4b.<br />

The calculated activation energies for the reaction<br />

between cellulose and <strong>AKD</strong> can be calculated from<br />

Fig 4b to be 72 kJ/mole at pH 4 to 46 kJ/mole at pH 10<br />

(Lindström and O´Brian 1986b). As expected the activa-<br />

Fig 4. (a) The quantity In a/(a-x) versus reaction time, t, where a is the maximum<br />

reacted <strong>AKD</strong> and x is the reacted amount of <strong>AKD</strong> at different pH-values. (b) The<br />

reaction rate constant K, from Eq 2, versus 1/T, where T is the absolute temperature.<br />

(Bleached softwood kraft pulp). (Lindström and O´Brian 1986b).<br />

Nordic Pulp and Paper Research Journal Vol 23 no. 2/2008 205


tion energy decreases for the nucleophilic reaction the<br />

higher the pH.<br />

Sizing accelerators<br />

The reaction between <strong>AKD</strong> and cellulose is, however<br />

slow and <strong>sizing</strong> accelerators are invariably used in commercial<br />

operations, whereby the reaction rate easily can<br />

be increased 20 times. The most important <strong>sizing</strong> accelerators<br />

are:<br />

· HCO-3<br />

· Basic polymers with amine groups<br />

The HCO -<br />

3-ion has a unique ability to catalyse the reaction<br />

between <strong>AKD</strong> and cellulose. HCO -<br />

3- -ions are often<br />

inherently present in natural systems, e.g. when CaCO3 is<br />

used as filler, but it is also a general practice to add<br />

NaHCO3 to increase the alkalinity of the stock.<br />

Fig 5a shows how the reaction of <strong>AKD</strong> in the presence<br />

of NaHCO3 is catalysed. The acceleration has been<br />

quantified using the above reaction rate expression. A<br />

further analysis of the so obtained reaction rate constant<br />

reveals that the reaction rate is proportional also to<br />

[HCO -<br />

3]. Hence, the reaction follows the equation:<br />

dx<br />

dt K<br />

-<br />

= [cellulose] ⋅[<strong>AKD</strong>] ⋅[HCO<br />

]<br />

0 3<br />

This equation clearly suggests that there is a trimolecular<br />

reaction between cellulose, <strong>AKD</strong> and HCO 3 taking place.<br />

The suggested mechanism of catalysis is given in Fig 5b.<br />

Polymeric amines (e.g. PAMAM-EPI resins) having<br />

amino groups with a free electron pair are classic <strong>sizing</strong><br />

accelerators for <strong>AKD</strong> (Lindström and Söderberg 1986d;<br />

Thorn et al. 1993; Cooper et al. 1995). Several different<br />

Fig 5. (a) The quantity ln a/(a-x) versus reaction time, t, where a is the maximum reacted <strong>AKD</strong> and x is the reacted<br />

amount of <strong>AKD</strong> at different bulk concentrations of NaHCO 3. (b) A suggested mechanism of catalysis with NaHCO 3.<br />

(Lindström and Söderberg 1986d).<br />

206 Nordic Pulp and Paper Research Journal Vol 23 no. 2/2008<br />

[5]<br />

Fig 6. Synergistic effects between HCO -<br />

3and PAMAM-EPI resin on the reaction<br />

rate constant, K, when simultaneously used in <strong>AKD</strong>-<strong>sizing</strong>. ❍ = <strong>sizing</strong> in deionized<br />

water at different pH-values. ●<strong>–</strong> = <strong>sizing</strong> in the presence of 0.1% PAMAM in deionized<br />

water (Lindström and Söderberg 1986a).<br />

types of condensation polymers have been investigated<br />

over the years and occur in commercial formulations. The<br />

polymeric <strong>sizing</strong> accelerators are often either added to the<br />

<strong>AKD</strong>-dispersion (”rapid curing dispersion”) or used separately<br />

as combined accelerator and retention aid. Moreover<br />

the effects of HCO -<br />

3 are synergistic, as shown in Fig 6.<br />

<strong>AKD</strong>-hydrolysis<br />

<strong>AKD</strong> can also react with water forming the β-keto acid,<br />

which spontaneously decarboxylates forming the corresponding<br />

ketone, as shown in Fig 1. <strong>AKD</strong> is, however,<br />

stable at room temperature at acidic pH-values, allowing<br />

storage at the time scale of months.<br />

It is well known that <strong>AKD</strong> can be<br />

the subject to alkaline hydrolysis. It<br />

is known that CaCO 3 may induce<br />

hydrolysis, particularly precipitated<br />

calcium carbonates (PCC) having<br />

higher pH-values due to residual<br />

alkali (Colasurdo and Thorn 1992;<br />

Novak and Rende 1993; Bottorff<br />

1994; Jiang and Deng 2000). As<br />

<strong>AKD</strong> is strongly adsorbed onto PCC<br />

and gronud calcium carbonate<br />

(GCC), it is advantageous to preadsorb<br />

cationic polymers onto carbonate<br />

fillers in order to block <strong>AKD</strong>deposition<br />

(Esser and Ettl 1997).<br />

There have, however, been few systematic<br />

and quantitative investigations<br />

in this field.<br />

In a recent investigation from this<br />

lab, the hydrolysis was studied using<br />

14 C-labelled <strong>AKD</strong> (Lindström and<br />

Glad-Nordmark 2007b). It was<br />

found that NaHCO 3 catalyzed the<br />

hydrolysis reaction, as shown in


Fig 7a. The mechanism is probably<br />

analogous to the mechanism of the<br />

catalysis of the cellulose-<strong>AKD</strong><br />

reaction shown in Fig 7b, because<br />

analysis of the reaction kinetics of the<br />

reaction pointed to a trimolecular<br />

reaction mechanism. In this investigation<br />

it was also found that divalent<br />

metal ions (Ca 2+ , Mg 2+ , Ba 2+ ) catalyzed<br />

the hydrolysis reaction (see Fig 8a),<br />

whereas the anion or monovalent electrolytes<br />

had no effects. It is most likely<br />

that the divalent cation functions<br />

as a Lewis type of catalyst (Schinzer<br />

1986), the mechanism of which is<br />

shown in Fig 8b. Hydrolysis only<br />

takes place at high temperatures and<br />

during standard laboratory forming<br />

and drying conditions hydrolysis is<br />

insignificant.<br />

The hydrolysis product of <strong>AKD</strong>, the<br />

ketone in Fig 1, has a slight positive<br />

effect on <strong>sizing</strong>, provided there is already<br />

some reacted <strong>AKD</strong> present in the<br />

paper (Lindström and Söderberg<br />

1986a).<br />

Amount of <strong>AKD</strong> required for <strong>sizing</strong><br />

The required amount of <strong>AKD</strong> for<br />

<strong>sizing</strong> for a given pulp depends on a number of factors<br />

and is also linked to a number of wet-end factors. Critical<br />

is the retention of the size and the extent of reaction<br />

together with the nature of the pulp furnish together with<br />

the structure of the sheet.<br />

Retention is critical, as recirculated size can be subject<br />

to hydrolysis. The extent of reaction depends on the<br />

drying conditions together with the presence of size<br />

accelerators. The extent of reaction for <strong>AKD</strong>-sizes is<br />

dependent on the fraction of fibre surface exposed to the<br />

air phase, because it is only size on free surfaces, which<br />

can spread and potentially react. <strong>AKD</strong> spreads all over<br />

the sheet and <strong>sizing</strong> with <strong>AKD</strong> is not dependent on size<br />

agglomeration, which is critical for instance with<br />

soap/alum <strong>sizing</strong>.<br />

The extent of reaction can be quite high under ideal<br />

laboratory conditions, but under practical mill conditions<br />

it is often in the range between 15-40%.<br />

In an early publication (Lindström and Söderberg 1986a)<br />

the amount required for <strong>sizing</strong> was investigated for<br />

various extractive free pulps. Defining the onset of full<br />

<strong>sizing</strong> as Cobb 60 = 25 g/m 2 , the required amount of <strong>AKD</strong><br />

necessary for <strong>sizing</strong> is directly proportional to the BET<br />

surface area of the papers as shown in Fig 9. By using<br />

surface balance measurements to determine the collapse<br />

value of the monolayer the planar oriented monolayer<br />

surface area of <strong>AKD</strong> can be calculated to 24 Å 2 per molecule.<br />

It is important to emphasize that <strong>sizing</strong> is uniquely<br />

defined by the reacted amount of <strong>AKD</strong> (Lindström and<br />

Söderberg 1986a; Johansson and Lindström 2004b).<br />

Using this surface area it can be calculated from Fig 9<br />

Fig 7. (a) The effect of NaHCO 3 on <strong>AKD</strong> hydrolysis, (b) Suggested hydrolysis mechanism.<br />

Fig 8. (a) The effect of CaCl 2 on <strong>AKD</strong> hydrolysis, (b) Suggested hydrolysis mechanism by Ca 2+ (Lewis acid catalysis).<br />

(Lindström and Glad-Nordmark, 2007b).<br />

Fig 9.The required amount of reacted <strong>AKD</strong>, necessary to obtain Cobb 60 = 25 g/m 2<br />

for various pulps (extractives free).<br />

that it is only necessary to cover 4% of the total surface<br />

area for a given pulp in order to obtain <strong>sizing</strong>. In practice<br />

the sweeping action of the fatty acid chains will, of course,<br />

cover a much larger area. Ström and co-workers<br />

(Ström et al. 1992) determined the required surface coverage<br />

to 15% using ESCA. The electrons from carbon<br />

atoms emitted in the ESCA analysis partly come from<br />

carbon atoms underneath the fatty acid layer, so the surface<br />

coverage values should not be directly compared.<br />

Nordic Pulp and Paper Research Journal Vol 23 no. 2/2008 207


Practical aspects and comparisons between different<br />

<strong>sizing</strong> agents<br />

In Table 2, some major aspects of different <strong>sizing</strong> agents<br />

have been compared. Rosin <strong>sizing</strong> is basically restricted<br />

to acidic pH-values and so both <strong>AKD</strong> and ASA are basically<br />

restricted to neutral/alkaline papermaking, although<br />

ASA may be used at slightly acidic pH-values (Roberts<br />

1997). Electrolytes are basically negative for all <strong>sizing</strong><br />

agents because they interfere with retention aid use and<br />

decrease their affinity to fibre surfaces. Divalent metal<br />

ions are devastating for rosin <strong>sizing</strong> because they compete<br />

with aluminium species in the complexation process<br />

and for <strong>AKD</strong> <strong>sizing</strong>, because they catalyze the <strong>AKD</strong><br />

hydrolysis reaction. Fines/fillers have a large surface area<br />

consuming the <strong>sizing</strong> agent. Fillers can generally not be<br />

sized because reactive sizes do not react with fillers and<br />

the aluminium resinate (rosin-aluminium complex)<br />

cannot be anchored to the filler.<br />

The hydrolysis product of ASA can complex with Ca 2+ -<br />

ions so it may in principle be able to use for slack <strong>sizing</strong><br />

of calcium carbonates (Roberts 1997).<br />

Dissolved anionic substances are in almost all cases<br />

detrimental to size retention. The charged groups on the<br />

fibres are necessary for rosin <strong>sizing</strong>, and the higher the<br />

carboxyl group content, the easier is is to size the pulp<br />

with rosin. For <strong>AKD</strong>/ASA sizes, the charged groups are<br />

in general beneficial for retention processes. Acidic<br />

extractives may in principle be used as <strong>sizing</strong> agents in<br />

the presence of alum and non-ionic extractives contribute<br />

only slightly to <strong>sizing</strong>. Extractives of the fatty acid types<br />

are detrimental to <strong>AKD</strong>-<strong>sizing</strong>, because they interfere<br />

with retention and with the <strong>AKD</strong>-reaction (Lindström<br />

and Söderberg 1986c; Lidén and Tollander 2004; Åvitsland<br />

et al. 2006), but have been found not to interfere<br />

with spreading (Mattsson et al. 2003). Extractives have in<br />

general a slight positive effect on ASA-<strong>sizing</strong> because<br />

aluminium salts are used in conjunction with ASA<strong>sizing</strong>.<br />

The <strong>AKD</strong>-hydrolysis product has a slight positive<br />

effect on <strong>AKD</strong>-<strong>sizing</strong>, but is detrimental for ASA-<strong>sizing</strong><br />

because the diacid is amphiphatic and will overturn in the<br />

presence of aqueous liquids in contact with the sized<br />

paper. Aluminium salts are, of course, necessary for rosin<br />

<strong>sizing</strong>, but interfere with <strong>AKD</strong>-<strong>sizing</strong>, if they contribute<br />

sufficient acidity to decrease the HCO -<br />

3 content of the<br />

water. <strong>AKD</strong> can also be used in conjunction with rosin<br />

<strong>sizing</strong> for liquid packaging (Walkden 1991), but there is<br />

no simple mechanistic explanation for this synergism.<br />

Table 2. Comparison between different <strong>sizing</strong> agents.<br />

Rosin <strong>AKD</strong> ASA<br />

pH 4.2-5.0 7-8.5 5-8.5<br />

Electrolytes - - - - -<br />

Fines/Fillers - - -<br />

Dissolved an. substances - - - - - -<br />

Fibre-COOH + + + +<br />

Extractives + - - (+)<br />

Hydrolysis products irrelevant (+) - -<br />

Aluminium sulfate + + +<br />

Stock temperature - - - -<br />

Lactic acid resistance - + -<br />

208 Nordic Pulp and Paper Research Journal Vol 23 no. 2/2008<br />

Stock temperature has a strongly negative effect, particularly<br />

on rosin soap <strong>sizing</strong>, but also on dispersion <strong>sizing</strong>. A<br />

higher temperature leads to aggregation of precipitated<br />

aluminum resinates and oxolation of aluminium species<br />

making them loose some of their cationic charge characteristics.<br />

For synthetic sizes a higher stock temperature<br />

leads to a higher rate of hydrolysis of non-retained size.<br />

Neither rosin sizes nor ASA-sizes can protect paper<br />

against liquids containing strongly coordinating species<br />

like lactic acid<br />

Acknowledgement:<br />

Veronica Sundling is acknowledged for editing the text.<br />

Literature<br />

Asakura, K., Iwamoto, M. and Isogai, A. (2006a): The effects of <strong>AKD</strong> oligomers<br />

present in <strong>AKD</strong> wax on dispersion stability and paper performance, Nord. Pulp<br />

Paper Res. J. 21(2), 245-252.<br />

Asakura, K., Iwamoto, M. and Isogai, A. (2006b): Influence of fatty acid anhydride<br />

components present in <strong>AKD</strong> wax on emulsion stability and paper <strong>sizing</strong> performance,<br />

APPITA, 59(4), 285-290.<br />

Bottorff, K.J. (1993): New insights into the <strong>AKD</strong> <strong>sizing</strong> mechanism, Nord. Pulp<br />

Paper Res. J. 8(1), 86-95.<br />

Bottorff, K.J. (1994): <strong>AKD</strong> <strong>sizing</strong> mechanism: A more definitive description, Tappi<br />

J. 77(4), 105-116.<br />

Brungardt, C.L. and Gast, J.C. (1996): Alkenyl-substituted <strong>sizing</strong> agents for precision<br />

converting grades of fine paper, Tappi Proc. 1996 Papermakers<br />

Conference, TAPPI Press, Atlanta, USA, 297-308.<br />

Brungardt, C.L. and Varnell, D.F. (2005): The effect of ketene dimer melting<br />

point on the rate of <strong>sizing</strong> development, In: Advances in Paper Science and<br />

Technology: 13th Fundamental Research Symposium, Cambridge, I´Anson, S.J.<br />

(ed.), Pulp and Paper Fundamental Research Society, Bury, UK, 193-209.<br />

Cazabat, A.M. (1989): The dynamics of wetting, Nord. Pulp Paper Res. J 4(2),<br />

146-154.<br />

Champ, S. and Ettl, R. (2004): The dynamics of alkylketene dimer (<strong>AKD</strong>) retention,<br />

J. Pulp Paper Sci. 30(2), 322-328.<br />

Colasurdo, A.R. and Thorn, I. (1992): The interaction of alkyl ketene dimer with<br />

other wet-end additives, Tappi J. 75(9), 205-211.<br />

Cooper, C., Dart, P., Nicholass, J. and Thorn, I. (1995): The role of polymers in<br />

<strong>AKD</strong> <strong>sizing</strong>, Paper Technol. (May), 30-34.<br />

Downey, W.F. (1949): Higher alkyl ketene dimer emulsion. US Pat. 2,627, 477.<br />

Dumas, R.W. (1975): An overview of cellulose reactive sizes, Tappi J. 64(1), 43.<br />

Eklund, D. and Lindström, T. (1991): Paper Chemistry <strong>–</strong> An introduction, DT<br />

Paper Science Publications, Grankulla, Finland.<br />

Esser, A. and Ettl, R. (1997): On the mechanism of <strong>sizing</strong> with alkyl ketene dimer<br />

(<strong>AKD</strong>): physico-chemical aspects of <strong>AKD</strong> retention and <strong>sizing</strong> efficiency,<br />

Fundamentals of Papermaking Materials: 11th Fund. Res. Symp. Cambridge,<br />

Fundamental Res. Comm. and Pira International, 997-1020.<br />

Garnier, G., Bertin, M. and Smrckova, M. (1999): Wetting dynamics of alkyl<br />

ketene dimer on cellulosic model surfaces, Langmuir, 15(22), 7863-7869.<br />

Garnier, G. and Godbout, L. (2000): Wetting behaviour of alkyl ketene dimer on<br />

cellulose and model surfaces, J. Pulp Paper Sci. 26(5), 194-199.<br />

Garnier, G., Wright, J., Godbout, L. and Yu, L. (1998): Wetting mechanism of<br />

alkyl ketene dimers on cellulose films, Coll. Surf. A: Phys. Chem. Eng. Asp. 145,<br />

153-165.<br />

Hardell, H.-L. and Woodbury, S. E. (2002): A new method for the analysis of <strong>AKD</strong><br />

oligomers in papermaking systems, Nord. Pulp Paper Res. J. 17(3), 340-335.<br />

Hasegawa, M., Isogai, A. and Onabe, F. (1997): Alkaline <strong>sizing</strong> with alkylketene<br />

dimers in the presence of chitosan salts, J. Pulp Paper Sci. 23(11), J528-J531.<br />

Hodgson, K.T. (1994): A <strong>review</strong> of paper <strong>sizing</strong> using alkyl ketene dimer vs. alkenyl<br />

succinic anhydride, APPITA, 47(5), 402.<br />

Horn, D. (2001): Exploring the nanoworld of interfaces and their functions


during paper manufacturing and upgrading, Wochenbl. für Papierfab.<br />

129(23/24), 1589-1596.<br />

Hubbe, M. ,A. (2006): Paper´s resistance to wetting <strong>–</strong> A <strong>review</strong> of internal <strong>sizing</strong><br />

chemicals and their effects, Bioresources, 2(1), 106-145.<br />

Hutton, B. and Chen, W. (2004): Sizing effects via <strong>AKD</strong> vaporisation, 58th Appita<br />

Annual Conf. and Exhibition, Canberra, ACT, Australia.<br />

Isogai, A. (1997a): Factors influencing on retention of alkyl ketene dimer, The<br />

Fundamentals of Papermaking, Trans. of the 11th Fund. Res. Symp. held at<br />

Cambridge, Pira International, 1047-1071.<br />

Isogai, A. (1997b): Effect of cationic polymer addition on retention of alkyl ketene<br />

dimer, J. Pulp Paper Sci. 23(6), J276-J281.<br />

Isogai, A. and Asakura, K. (2001a): Paper <strong>sizing</strong> by liquid <strong>–</strong> type of ketene<br />

dimers Part 1. Comparison with alkyl ketene dimer and oleic anhydride, Nord.<br />

Pulp Paper Res. J. 16(2), 103-107.<br />

Isogai, A. and Asakura, K. (2001b): Paper <strong>sizing</strong> by liquid-type ketene dimers.<br />

Part 2. Effects of drying methods, UV and air-blowing treatments on PCC-filled<br />

sheets, Nord. Pulp Paper Res. J. 16(2), 108-112.<br />

Isogai, A., Kitaoka, C. and Onabe, F. (1997): Effects of carboxyl groups in pulp<br />

on retention of alkyl ketene dimer, J. Pulp Paper Sci. 23(5), J217-J219.<br />

Jiang, H. and Deng, Y. (2000): The effects of inorganic salts and precipitated<br />

calcium carbonate filler on the hydrolysis kinetics of alkyl ketene dimer, J. Pulp<br />

Paper Sci. 26(6), 208-213.<br />

Johansson, J. and Lindström, T. (2004b): A study on <strong>AKD</strong>-size retention, reaction<br />

and <strong>sizing</strong> efficiency. Part 2: The effects of electrolytes, retention aids, shear<br />

forces and mode of addition on <strong>AKD</strong>-<strong>sizing</strong> using anionic and cationic <strong>AKD</strong>-dispersions,<br />

Nord. Pulp Paper Res. J. 19(3), 336-344.<br />

Lidén, J. and Tollander, M. (2004): Extractives in totally chlorine free bleached<br />

birch pulp and their effect on alkyl ketene dimers and alkeny succinic anhydride<br />

sizes, Nord. Pulp Paper Res. J. 19(4), 466-469.<br />

Lindström, T. and Glad-Nordmark, G. (2007a): A study of <strong>AKD</strong>-size retention,<br />

reaction and <strong>sizing</strong> efficiency. Part 3: The effects of fibre charge density and electrolyte<br />

concentration on size retention, Nord. Pulp Paper Res. J. 22(2), 161-166.<br />

Lindström, T. and Glad-Nordmark, G. (2007b): A study of <strong>AKD</strong>-size retention,<br />

reaction and <strong>sizing</strong> efficiency. Part 4: Mechanism of <strong>AKD</strong>-hydrolysis, Nord. Pulp<br />

Paper Res. J. 22(2), 167-171.<br />

Lindström, T. and O´Brian, H. (1986b): On the mechanism of <strong>sizing</strong> with alkyl<br />

ketene dimers.Part II. The kinetics of reaction between alkyl ketene dimers and<br />

cellulose, Nord. Pulp Paper Res. J. 1(1), 34-42.<br />

Lindström, T. and Söderberg, G. (1986a): On the mechanism of <strong>sizing</strong> with alkyl<br />

ketene dimers. Part I. Studies on the amount of alkyl ketene dimer required for<br />

<strong>sizing</strong> of different pulps, Nord. Pulp Paper Res. J. 1(1), 26-33.<br />

Lindström, T. and Söderberg, G. (1986c): On the mechanism of <strong>sizing</strong> with alkyl<br />

ketene dimers. Part III. The role of pH, electrolytes, retention aids, extractives, Calignosulfonates<br />

and mode of addition on alkyl ketene dimer retention, Nord. Pulp<br />

Paper Res. J, 1(2), 31-38.<br />

Lindström, T. and Söderberg, G. (1986d): On the mechanism of <strong>sizing</strong> with alkyl<br />

ketene dimers. Part IV. The effects of HCO 3-ions and polymeric reaction accelerators<br />

on the rate of reaction between alkyl ketene dimers and cellulose, Nord. Pulp<br />

Paper Res. J. 1(2), 39-45.<br />

Lundqvist, Å. and Ödberg, L. (1997): Surface energy characterization of surface<br />

modified cellulosic fibres by inverse gas chromatography, in The Fundamentals of<br />

Papermaking Materials. Trans of the 11th Fund. Res. Symp. Sept. 1997, held at<br />

Cambridge, UK, Pira International, Leatherhead, Surrey, UK, Vol 1, pp. 751-769.<br />

Luner, P.E. and Oh, E. (2001): Characterization of the surface free energy of cellulose<br />

ether films, Coll. Surf. A. Phys. Chem. Eng Asp. 181, 31-48.<br />

Mattsson, R., Lindström, D., Sterte, J. and Ödberg, L. (2003): Influence of abietic<br />

acid, betulinol, sodium oleate and tripalmitine on the migration of <strong>AKD</strong> in<br />

paper, J. Pulp Paper Sci. 29(8), 281-285.<br />

Mattsson, R., Sterte, J. and Ödberg, L. (2001): Colloidal stability of alkyl ketene<br />

dimer (<strong>AKD</strong>) dispersion: Influence of shear, electrolyte concentration, polyelectrolytes<br />

and surfactants, The Science of Papermaking, 12th Fundamental Res.<br />

Symp. Oxford, The Pulp and Paper Fundamental Res. Soc., 393-413.<br />

Mattsson, R., Sterte, J. and Ödberg, L. (2002): Sizing with pre-flocculated alkyl<br />

ketene dimer (<strong>AKD</strong>) dispersions, Nord. Pulp Paper Res. J. 17(3), 240-245.<br />

Nahm, S. (1986): Direct evidence for covalent bonding between ketene dimer<br />

<strong>sizing</strong> agents and cellulose, J. Wood Chem. Technol. 6(1), 89-112.<br />

Novak, R. and Rende, D. (1993): Size reversion in alkaline papermaking, Tappi J.<br />

76(8), 117-120.<br />

Petander, L., Ahlskog, T. and Juppo, A. J. (1998): Strategies to reduce <strong>AKD</strong><br />

deposits on paper machines, Paperi Puu, 80(2), 100-103.<br />

Reynolds, W. F. (1989): The Sizing of Paper, Tappi Press, Atlanta, GA, USA.<br />

Riebeling, U., Jeurissen, H.F.M., De Clercq, A. and Prinz, M. (1996): Ein neues<br />

Masseleimungskonzept-Wirkungsverbesserung von <strong>AKD</strong> durch Einsatz hydrophober<br />

amphoterer Polymere, Wochenbl. für Papierfab. 22, 997-1002.<br />

Riebeling, U., Jeurissen, H.F.M., Höhr, L. and Schall, N. (1999):<br />

Wirkungsverbesserung von <strong>AKD</strong> durch Einsatz hydrophober amphoterer Polymere<br />

(Teil 2), Wochenbl. für Papierfab. 5, 318-322.<br />

Roberts, J.C. (1997): A <strong>review</strong> of advances in internal <strong>sizing</strong> of paper, The<br />

Fundamentals of Papermaking, Trans. of the 11th Fund. Res. Symp., Cambridge,<br />

Pira International, 209-263.<br />

Roberts, J.C. and Garner, D.N. (1985): The mechanism of alkyl ketene dimer<br />

<strong>sizing</strong> of paper. Part 1, Tappi J. 68(4), 118-121.<br />

Roberts, J.C., Garner, D.N. and Akpabio, U.D. (1985): Neutral <strong>sizing</strong> and the<br />

mechanism of alkyl ketene dimer <strong>sizing</strong>, “Papermaking raw materials” Trans. of<br />

the eighth Fundamental Res. Symp. held at Oxford, 1985, 2 (3), 815-837.<br />

Schinzer, D. (1986): Selectives in Lewis Acid Promoted Reactions. Kluwer<br />

Academic Publishers, Dordrecht, The Netherlands<br />

Seppänen, R., Tiberg, F. and Valignat, M.-P. (2000): Mechanism of internal<br />

<strong>sizing</strong> by alkyl ketene dimers (<strong>AKD</strong>): The role of the spreading manolayer precourser<br />

and autophobicity, Nord. Pulp Paper Res. J. 15(5), 452-458.<br />

Shchukarev, A.V., Mattsson, R. and Ödberg, L. (2003): XPS imaging of surface<br />

diffusion of alkyl ketene dimer on paper surfaces, Coll. Surf. A: Phys. Chem. Eng.<br />

Asp. 219(1-3), 35-43.<br />

Shen, W., Brack, N., Ly, H., Parker, I., Pigram, P.J. and Liesegang, J. (2001a):<br />

Mechanism of <strong>AKD</strong> migration studied on glass surfaces, Coll. Surf. A: Phys.<br />

Chem. Eng. Asp. 176(2-3), 129-137.<br />

Shen, W. and Parker, I. (2003): An experimental investigation of the redistribution<br />

behaviour of alkyl ketene dimers and their correponding ketones, Coll. Surf.<br />

A: Phys. Chem. Eng. Asp. 212(2-3), 197-209.<br />

Shen, W. and Parker, I.(2003): A study of non-solid behaviour of <strong>AKD</strong> wax,<br />

APPITA, 56(6), 442-444.<br />

Shen, W., Parker, I., Brack, N. and Pigram, P.J. (2001b): A simplified approach<br />

to understanding the mechanism of <strong>AKD</strong> <strong>sizing</strong>, APPITA, 54(4), 352-356.<br />

Shen, W., Zhang, H. and Ettl, R. (2005): Chemical composition of “<strong>AKD</strong> vapour”<br />

and its implication to <strong>AKD</strong> vapour <strong>sizing</strong>, Cellulose, 12(6), 641-652.<br />

Ström, G., Carlsson, G. and Kiær, M. (1992): Bestimmung der <strong>Alkyl</strong>ketenedimerverteilung<br />

in Probeblättern mittels elektron spektroskopie (ESCA), Wochenbl.<br />

Papierfabr. 120(15), 606-611.<br />

Thorn, I., Dart, P. ,J. and Main, S. ,D. (1993): The use of cure promotors in alkaline<br />

<strong>sizing</strong>, Paper Technol. 34(1), 41-45.<br />

Walkden, S. ,A. (1991): Alkaline advance helps liquid packaging board meet<br />

rigorous specifications, Tappi J. 74(4), 103-104.<br />

Yu, L. and Garnier, G. (2002): The role of vapour deposition during internal<br />

<strong>sizing</strong>: A comparative study between ASA and <strong>AKD</strong>, J. Pulp Paper Sci. 28(10),<br />

327-331.<br />

Åvitsland, G., Sterner, M., Wågberg, L. and Ödberg, L. (2006): <strong>AKD</strong> <strong>sizing</strong> TCF<br />

and ECF bleached birch pulp characterized by peroxide edge wicking index, Nord.<br />

Pulp Paper Res. J. 21(2), 237-244.<br />

Ödberg, L., Lindström, T., Liedberg, B. and Gustavsson, J. (1987): Evidence<br />

for β-ketoester formation during the <strong>sizing</strong> of paper with alkyl ketene dimers,<br />

Tappi J. 70(4), 135-139.<br />

Manuscript received October 8, 2007<br />

Accepted March 16, 2008<br />

Nordic Pulp and Paper Research Journal Vol 23 no. 2/2008 209

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!